Issue
EPJ Appl. Metamat.
Volume 5, 2018
Terahertz metamaterials
Article Number 10
Number of page(s) 9
DOI https://doi.org/10.1051/epjam/2018006
Published online 24 October 2018

© E. Akmansoy and S. Marcellin, published by EDP Sciences, 2018

Licence Creative CommonsThis is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

1 Introduction

All-dielectric metamaterials (ADMs) are the promising “inflection” of metamaterials to go beyond their limits. ADMs are an alternative to metallic metamaterials. The advantages of ADMs come from their low losses and their simple geometry: they do not suffer from Ohmic losses and thus they may benefit from low energy dissipation, specially, ceramics of high dielectric permittivity and high quality factor [1,2]. In the microwave, their quality factor is greater than that of metallic metamaterials [3], and it is consequently also the case in terahertz (THz) and optical frequencies. ADMs are partly inspired by the work of Richtmyer who developed the theory of dielectric resonators, which is based on the fact that “the dielectric has the effect of causing the electromagnetic field [. . .] to be confined to the cylinder itself and the immediately surrounding region of space” [4]. Taking the matter further, O'Brien and Pendry opened the way for ADMs by considering the periodic lattice of high permittivity resonators (HPRs), thus demonstrating artificial magnetism in the microwave [2, 5]. ADMs rely on the first two modes of the Mie resonances of HPR. The first mode results in resonant effective permeability that can have negative values, while the second one results in resonant effective permittivity that can also have negative values. When both are simultaneously negative, the ADM is called “double negative” (DNG) and its effective refractive index is then negative [611]. The unit cell of ADM thus comprises two subwavelength building blocks of simple geometry [12,13]. In analogy with chemical molecules, the unit cell is generally called a meta-dimer [14], and the two building blocks are called meta-atoms. As the two are different, the unit cell is a hetero-dimer.

The large applications of ADMs (for a review, see Ref. [15]) include perfect reflectors [16], perfect absorbers [17, 18], zero-index metamaterials [19], optical magnetic mirror [20], and Fano resonances [21, 22]. ADMs have been demonstrated from the microwave to the optical domain. Artificial magnetism [3, 2325] and negative effective refractive index [26, 27] have been, theoretically and experi­mentally, shown in the GHz regime. Even though artificial magnetism provided by ADMs has been experimentally demonstrated in the THz [28] and infrared ranges [2932], DNG refractive index has not been yet demonstrated, which impedes ADMs to be the equivalent of their metallic counterpart. Besides, THz radiation is widely defined as electromagnetic radiation in the frequency range 0.3–3 THz. It permits the obtention of physical data that are not accessible using X-ray or infrared radiation and it thus finds many applications in imaging, security, quality inspection, chemical sensing, astronomy, etc. On their part, metamaterials have evolved towards the implementation of photonic components [33]. HPRs are well suited for metamaterial applications in the low THz frequency range [34].

In the following, we report on the mode coupling effect which plays a dominant role in the electromagnetic properties of metamaterials [3539], notably, in the achievement of a negative effective index. Magnetic and electric mode coupling effects in ADMs have been separately studied in the microwave by Zhang et al. [36]. Herein, we report on the magneto-electric mode coupling effects in the THz domain, and we show that negative effective refractive index requires sufficiently strong mode coupling [40]. We also show that adjusting the mode coupling allows us to attain near-zero values of the refractive index, or even null effective index. Moreover, we highlight that the strongest values of the mode coupling lead to frequency mode degeneracy, for which the refractive index is undetermined.

2 Simulations

We consider a 2D ADM whose unit cell consists of two distinct building blocks, a magnetic block and an electric one, the former resonating in the first mode of Mie resonances and the latter resonating in the second mode [3, 13, 27, 41]. The first mode is thus referred to as the magnetic mode and the second one as the electric mode. The ADM consists of one infinite layer made up of two sets of high permittivity square cross-section dielectric cylinders that are perpendicular to the incident wave vector (Fig. 1). The two sets of infinitely long HPRs are actually interleaved. The incident polarization is transverse electric (TE), i.e., the electric field is perpendicular to the axis of the cylinders. We studiy both spatial mode coupling and frequency mode coupling. The ADM has been numerically simulated by the means of the finite elements method software HFSSTM, which yields the S-parameters. The side lengths of the resonators are initially am = 60 μm for the magnetic mode and ae = 90 μm for the electric one, while the lattice period is lp = 260 μm. The HPRs are equidistant and therefore, the distance between two of them is half the lattice period p2 = lp/2 = 130 μm. The relative permittivity of the dielectric is εr =94 (titanium dioxide − TiO2) and the loss tangent increases between tan δ = 0.009 and 0.015 in the considered frequency range [42, 43]. We are thus dealing with a high refractive index bulk material (NTiO2⋍10).

thumbnail Fig. 1

Schematic layout of the 2D ADM (cross-sectional view). The ADM consists of one infinite layer along the x direction and is made up of two interleaved sets of high permittivity square cross-section dielectric cylinders, which resonate in the first two modes of Mie resonances: the small set resonates in the magnetic mode and the second one in the electric mode. The equidistant cylinders are infinite along the y direction and their side lengths are am = 60 μm and ae = 90 μm, respectively. Their relative permittivity is er = 94. The unit cell actually consists of two subwavelength distinct building blocks. The lattice period is lp= 260 μm and p2 is half the lattice period. The incident wave is transverse electric (TE).

3 Results and discussion

3.1 Negative index and mode coupling

We study the mode coupling between the first two modes of Mie resonances depending on the lattice period lp (spatial mode coupling), and then depending on the frequency overlapping of the two modes (frequency mode coupling). The results of the simulation, namely the minimum neffmin of the real part of the effective index neff as a function of the lattice period and as a function of the frequency spacing between the two modes, are reported in Figures 2 and 3, respectively. They show that the mode coupling should be strong enough to ensure negative effective index. The minima of the real part of the effective index neff are neffmin = −2.2 and −1.9, respectively. On the one hand, increasing the mode coupling is obtained by decreasing the lattice period lp. On the other hand, it is obtained by decreasing the frequency of the second mode of Mie resonances, which stems from the increasing of the side length ae of the electric resonator. When the mode coupling is sufficient, the bandwidth of the negative effective index neff increases with it (see insets in Figs. 2 and 3). The frequency range of negative effective index is given by the relation [44] (1)

where ' and ” respectively denote the real and imaginary parts of the permeability μ and the permittivity ε.

thumbnail Fig. 2

Spatial mode coupling: minimal value of the effective refractive index neff as a function of the distance p2 between the two resonators. The side lengths of the resonators are am = 60 μm and ae = 90 μm, respectively. Insets: real (purple) and imaginary part (green) of the effective index neff; the shaded area denotes the bandwidth of negative effective index.

thumbnail Fig. 3

Frequency mode coupling: minimal value of the effective refractive index neff as a function of the side length ae of the electric resonator, namely, the frequency of the electric mode is varying. Insets: real (purple) and imaginary part of (green) the effective index neff; the shaded area denotes the bandwidth of negative effective index. f11 and f12 are the resonances frequencies of the individual resonators (see the text).

3.2 Monomode coupling

To carry out our study, we first explored monomode coupling, that is, the mode coupling due to only one mode, which arises inside a layer whose unit cell only consists of one building block, the magnetic one or the electric one. We study both cases. Consequently, we only consider the spatial mode coupling and only varied the lattice period lp which is equal to the distance p1 between two resonators, lp = p1. Operating in the same frequency range, the side length of the magnetic resonator is am = 60 μm, while that of the electric resonator is ae = 90 μm. The results of the simulation are reported in Figures 4 and 5, for both cases, and they show that the two modes behave differently. Their respective frequencies (fmr, fer) are given by the minima of the S12 parameter [36]. The frequency fmr of the magnetic mode increases with the lattice period lp, whereas the frequency fer of the electric one decreases. These results are consistent with that of Zhang et al [36]. Both maxima of the absorption Amr and Aer (A(ω) = 1 − | S12 (ω) | 2− | S11(ω)|2) and the frequency of the resonance modes of the individual resonator are also reported. The latter provides a series of resonances whose frequency is determined by Cohn's model [1, 26, 45, 46]. (2)

where εr is the relative permittivity of the resonator, m and n the integers, a and b the side lengths of the resonator, and c is the velocity of light and its accuracy is about 5%. Equation 2 was used to design all the reported structures. For a square cross-section cylinder, a = b, and the frequencies of the first two modes of the individual resonator respectively correspond to m = n= 1 and m = 1 and n = 2. To exhibit negative refractive index, the involved resonance modes are the first mode of the magnetic resonator and the second mode of the electric resonator. Their frequencies are respectively f11 = 0.364 THz (a = am = 60 μm) and f12 = 0.384 THz (a = ae = 90 μm) [27]. These are obviously constant, while the maxima of the absorption are nearly constant. It can be noticed that, as the mode coupling increases, the distance between the resonance frequencies (fmr, fer) of the resonator inside the layer and that (f11, f12) of the individual resonator respectively increases (cf. Figs. 4 and 5, respectively), thus demonstrating the mode coupling.

thumbnail Fig. 4

Magnetic mode coupling: frequency (fmr) of the first mode of Mie resonances and of the maximum of absorption (Amr) as a function of the distance p1 between two resonators. The unit cell only comprises the magnetic block. The side length of the resonator is am = 60 μm. The dashed line denotes the resonance frequency (f11) of the mode of resonances of the individual resonator.

thumbnail Fig. 5

Electric mode coupling: frequency (fer) of the second mode of Mie resonances and of the maximum of absorption (Aer) as a function of the distance p1 between two resonators. The unit cell only comprises the electric block. The side length of the resonator is ae = 90 μm. The dashed line denotes the resonance frequency (f12) of the mode of resonances of the individual resonator.

3.3 Bimodal coupling

We next explore the mode coupling inside the ADM, namely, the unit cell now consists of the two building blocks. The variation of the resonance frequencies (fmr, fer) of both modes is reported in Figures 6 and 7 corresponding to the spatial mode coupling and the frequency mode coupling, respectively. These frequencies are still given by the minima of the S12 parameter [36]. Decreasing the lattice period lp, i.e., the distance p2 between the resonators, increases the mode coupling. Varying the overlapping of the two modes stems from the decreasing of the frequency of the electric mode, which also increases the mode coupling [35]. The curves are shaped as “tuning forks” and show that the frequencies (fmr, fer) of the two modes of resonance are moving closer together as the mode coupling increases. To highlight this effect, the frequencies (f11, f12) of the resonance modes of the individual resonator are again shown in these figures. For both the magnetic mode and the electric one, the distance between the frequency (fmr,fer) of the resonator inside the ADM and the frequency (f11, f12) of the individual one respectively increases as the mode coupling increases. This anew evidences the mode coupling inside the ADM. Theses curves also show that further increasing the mode coupling gives rise to a frequency degeneracy in both cases of mode coupling, that is, the two resonance frequencies of the two modes become equal, fmr = fer.

The S12 parameter and the absorption A are reported in Figure 8 as a function of frequency for several values of lattice period lp, which is relative to the spatial mode coupling. Two ranges of lattice period lp are considered, one is out of the frequency degeneracy regime (250 ≤ lp ≤ 400 μm) and the second one in the frequency degeneracy regime (200 ≤ lp ≤ 248 μm). It can be noticed that the frequency of both maxima of absorption are practically constant, whereas the frequency of both minima of the S12 parameter vary with the lattice period lp. The latter move closer together as the lattice period lp increases, until they merged, which corresponds to the frequency degeneracy. The crossing point is reached when the lattice period is equal to lpc = 250 μm. The merged minima of the S12 parameter are very weak (≲ − 51dB) for lp = 248 μm, which corresponds to the minimum of the real part of the effective index neffmin(lp = 248 μm) = −2.2. (See effective index curves inserted in Figs. 2 and 3 which are continuous, but at the limit of continuity.) For greater values of the mode coupling, that is, for lattice period lp smaller than 248 μm, the effective parameters could not be extracted, being not continuous through all frequencies [47]; the effective refractive index neff is then undetermined. We observed the same frequency degeneracy behavior when studying the frequency mode coupling (results not shown).

The mode coupling effect we report here is different from hybridization [48] which is observed with plasmonic metameterial [14, 49], split-ring resonators metametarials [50], inductor-capacitor resonators [51], cut wires [52], nano-wires [53], nano-rings [54], nano-particle dimers [55], or silicon nanoparticles [56]. In these cases, the coupling between the two identical meta-atoms which constitute the dimer leads from a trapped mode to the formation of new hybridized modes because it lifts the degeneracy of the mode of the individual meta-atoms. Hybridization may be used to yield negative refractive index [50, 51, 54]. The unit cell is then a homo-dimer and the negative effective index is achieved by playing with the mode coupling so as to overlap two hybridized modes which are of different kind: magnetic orelectric. In the case we report here, the mechanism is different since it takes the reverse way, because increasing the mode coupling leads from two separate modes to a trapped one. The unit cell of our ADM is a hetero-dimer because the two meta-atoms are not identical, and moreover, each one resonates in a different kind of mode: the magnetic one and the electric one. Hence, increasing the mode coupling leads to the trapped mode putting together the two separate modes. Nevertheless, the mode coupling has to be strong enough to ensure a negative effective index, the unit cell being either a homo-dimer or a hetero-dimer.

thumbnail Fig. 6

Spatial mode coupling: frequency of the first two modes of Mie resonances as a function of the distance p2 between two resonators which is half the lattice period lp. Blue and green colors denote the magnetic and the electric modes, respectively. Square dots and crossed dots denote the maxima of absorption (Amr, Aer) and the minimum of the S12 parameter (fmr, fer), respectively. The shaded area corresponds to negative value of the effective index neff. The side lengths of both resonators are am = 60 μm and ae = 90 μm, respectively. The dashed lines denote the frequencies of resonances (f11, f12) of the modes of the individual resonator.

thumbnail Fig. 7

Frequency mode coupling: frequency of the first two modes of Mie resonances (fmr, fer) as a function of the side length ae of the electric resonator, namely, the frequency of the electric mode is varying. The convention is the same as in Figure 6. The side length of the magnetic resonator is am = 60 μm and the lattice period lp is 260 μm.

thumbnail Fig. 8

Effect of spatial mode coupling: S12 parameter (log. scale (dB)) and absorption A (linear scale (lin.)) for several values of distance p2 between two resonators. The side lengths of the resonators are am = 60 μm and ae = 90 μm. (a) Out of frequency degeneracy regime (125 ≤ p2 ≤ 200 μm); (b) in frequency degeneracy regime (100 ≤ p2 ≤ 124 μm). The merged minima of the S12 parameter attain very weak values (≲ − 51 dB) when p2 = p2c = 124 μm, which correspond to the minimum of the refractive index neffmin = −2.2.

3.4 Effective parameters

In our ADM, a magnetic moment ensues from the magnetic mode giving rise to resonant effective permeability μeff. Similarly, an electric moment ensues from the electric mode giving rise to resonant effective permittivity eeff. The two are perpendicular to each other (see e.g., Fig. 1 in reference [41]). The mode coupling arises from the interaction between these electromagnetic moments and it changes with the separating distance. The side lengths of the resonators are afresh am = 60 μm and ae = 90 μm. Both resonances consequently modify the effective refractive index neff(ω) of the ADM. The effective electromagnetic parameters μeff, eeff and neff are reported in Figure 9 relative to two values of the lattice period lp = 360 μm and lp = 260 μm. They are extracted from the S-parameters using the common retrieval method described in references [47, 5763]. The antiresonance behavior of the effective permittivity eeff around the magnetic mode frequency, which is inherent in metamaterials, can be observed [6265]. In the former case (low mode coupling), the real part of the effective index neff does not reach negative values, whereas it does in the latter case (strong mode coupling), then satisfying equation 1. The minimum value of the real part of the effective index is then neffmin(lp =260 μm) = −1.5. In the former case, the real part of the effective index neff is below unity and close to zero. Its minimal value is neffmin(lp = 360 μm) ≲ 0.04, demonstrating that adjusting the mode coupling makes ADMs suitable for epsilon-near-zero (ENZ) metamaterials [66, 67] (see also Fig. 2), or even null effective index [19]. It can also be noticed that the mode coupling strongly enhances the amplitude of both resonances, notably the electric one, and that it brings closer together the two modes. In addition, as we are dealing with a high refractive index bulk material (NTiO2 ≃ 10), the wavelength inside it is about one-tenth of that in vacuum. Consequently, we calculated the static effective permittivity, i.e., beyond the resonances, from the Maxwell–Garnett model [68]. It is equal to eeff = 2.1 and eeff = 2.9 corresponding to both values of the lattice period lp = 360 μm and lp = 260 μm, respectively, which is in good agreement with the results of the simulation (see Fig. 9a and b-middle).

To engineer the electromagnetic properties of an ADM, one can consequently play with either mode couplings: the spatial mode coupling or the frequency mode cou­pling. Figure 10 gathers the role of both mode couplings: the stronger the two mode couplings, the more negative the effective index and the larger the bandwidth of negative effective index (cf. Figs. 2 and 3). Combining the two mode couplings, negative effective index as low as neff = −2.8 is obtained.

thumbnail Fig. 9

Effect of spatial mode coupling: effective electromagnetic parameters (real and imaginary parts): permeability μeff (top), permittivity eeff (middle) and effective index neff (bottom). (a) Low mode coupling: lattice period lp = 360 μm: the minimum value of the effective refractive index is neffmin < 0.04; (b) strong mode coupling: lattice period lp = 260 μm. The shaded area denotes the bandwidth of the negative effective index: the minimum value of the effective refractive index is neffmin = −1.5. The static effective permittivity values yielded by the Maxwell–Garnett formula, eeff = 2.1 and eeff = 2.9, are in good agreement with the result of the simulation.

thumbnail Fig. 10

Effect of both mode couplings (spatial mode coupling (SMC) and frequency mode coupling (FMC)): minimal value of the effective refractive index neff as a function of the distance p2 between two resonators for several values of the side length ae of the electric resonator. The side length of the magnetic resonator is am = 60 μm. The two arrows denote increasing mode coupling. The minimum of the real part of the effective index is neffmin = −2.8.

3.5 Dielectric function, phonons and strontium titanate (SrTiO3)

Other high permittivity materials, having low losses, can be investigated to study the mode coupling inside ADMs at THz frequencies, e.g., SrTiO3 [6972]. However, the dielectric function εr(ω) of these high permittivity materials is dispersive, because of lattice vibrations, namely, due to optical phonons [73]. Their frequency is in the THz range, and we are concerned with the transverse optical phonon of lowest frequency (TO1). The TO1 phonon frequency of SrTiO3 is 2.70 THz [69], while that of TiO2 is 5.6 THz [69, 73]. The dielectric function εr(ω) is described by the classical oscillator model or the Four-Parameter Semi Quantum (FPSQ) model [69, 71, 73]. Measurements in the THz frequency, reported in the literature, are in good agreement with these models for TiO2 [42, 43] and SrTiO3 [71, 72, 74, 75]. The two yield the same dielectric function εr(ω) at the operating frequency. The latter also reflect that the imaginary part of the dielectric function εr(ω) strongly increases around the frequency of the TO1 phonon. We also simulated a similar ADM but consisting of HPRs made of SrTiO3 and found that it exhibits the same mode coupling effects (results not shown here). The real part of the effective index reaches values as low as neffmin = −7, meaning that the effect of the mode coupling is stronger when a higher permittivity material is used (εr ≃ 300 [72] as compared with εr = 94), because this heightens the resonances.

However, losses (imaginary part of εr(ω)) resulting from the TO1 phonon greatly increase and therefore limit the operating range at THz frequencies. In addition, the real part of the dielectric function εr(ω) falls down at higher frequency [72, 74], which drastically modifies the Mie resonances. The lower permittivity of TiO2 leads to greater side lengths am and ae of each resonator (see Eq. (2)) and therefore, it facilitates their fabrication. Consequently, TiO2 is more suitable for ADM applications at THz frequencies.

4 Conclusion

We have studied mode coupling effects in ADMs at THz frequencies, and we show that the mode coupling has to be sufficiently strong to ensure negative effective index of refraction. Tuning the first two modes of Mie resonances of an ADM by adjusting the mode coupling allows the setting of the effective index from a near-zero value to a negative value. We studied both spatial mode coupling and frequency mode coupling. Increasing both brings the modes closer together until they are merged. Thus, we highlight the frequency degeneracy of the first two resonance modes, namely, the two frequencies are equal, and the effective index is then undetermined. At the crossing point, the effective index reaches its lowest value.

Author contribution statement

E.A. supervised the study, analyzed the results and wrote the paper. S.M. made the simulation and analyzed the results.

Acknowledgments

The authors thank Aloyse Degiron for helpful discussion. E.A. thanks Fabrice Rossignol and Jean-Pierre Ganne for their help. This work has been funded by the Agence Nationale de la Recherche TeraMetaDiel project (number ANR-12-BS03-0009).

References

  1. D. Kajfez, P. Guillon, Dielectric Resonators (Artech House, Dedham, MA, 1986) [Google Scholar]
  2. S. Fiedziuszko, I. Hunter, T. Itoh, Y. Kobayashi, T. Nishikawa, S. Stitzer, K. Wakino, Dielectric materials, devices, and circuits, IEEE Trans. Microw. Theory Tech. 50, 706 (2002) [CrossRef] [Google Scholar]
  3. B.-I. Popa, S.A. Cummer, Compact dielectric particles as a building block for low-loss magnetic metamaterials, Phys. Rev. Lett. 100, 207401 (2008) [CrossRef] [Google Scholar]
  4. R.D. Richtmyer, Dielectric resonators, J. Appl. Phys. 10, 391 (1939) [CrossRef] [Google Scholar]
  5. S. O’Brien, J.B. Pendry, Photonic band-gap effects and magnetic activity in dielectric composites, J. Phys.: Condens. Matter 14, 15 (2002) [Google Scholar]
  6. J.B. Pendry, D.R. Smith, Reversing light with negative refraction, Phys. Today 57, 37 (2004) [CrossRef] [Google Scholar]
  7. W.J. Padilla, D.N. Basov, D.R. Smith, Negative refractive index metamaterials, Mater. Today 9, 28 (2006) [CrossRef] [Google Scholar]
  8. D.R. Smith, W.J. Padilla, D.C. Vier, S.C. Nemat-Nasser, S. Schultz, Composite medium with simultaneously negative permeability and permittivity, Phys. Rev. Lett. 84, 4184 (2000) [CrossRef] [PubMed] [Google Scholar]
  9. R.W. Ziolkowski, Design, fabrication, and testing of double negative metamaterials, IEEE Trans Antenn Propag. 51, 1516 (2003) [CrossRef] [Google Scholar]
  10. N. Engheta, R.W. Ziolkowski, A positive future for double-negative metamaterials, IEEE Trans. Microw. Theory Tech. 53, 1535 (2005) [CrossRef] [Google Scholar]
  11. R. Liu, A. Degiron, J.J. Mock, D.R. Smith, Negative index material composed of electric and magnetic resonators, Appl. Phys. Lett. 90, 263504 (2007) [CrossRef] [Google Scholar]
  12. A. Ahmadi, H. Mosallaei, Physical configuration and performance modeling of all-dielectric metamaterials, Phys. Rev. B 77, 045104 (2008) [CrossRef] [Google Scholar]
  13. Q. Zhao, J. Zhou, F. Zhang, D. Lippens, Mie resonance-based dielectric metamaterials, Mater. Today 12, 60 (2009) [CrossRef] [Google Scholar]
  14. N. Liu, H. Liu, S. Zhu, H. Giessen, Stereometamate-rials, Nat. Photonics 3, 157 (2009) [CrossRef] [Google Scholar]
  15. S. Jahani, Z. Jacob, All-dielectric metamaterials, Nat Nano 11, 23 (2016) [Google Scholar]
  16. P. Moitra, B.A. Slovick, W. Li, I.I. Kravchencko, D.P. Briggs, S. Krishnamurthy, J. Valentine, Large-scale all-dielectric metamaterial perfect reflectors, ACS Photonics 2, 692 (2015) [CrossRef] [Google Scholar]
  17. X. Liu, Q. Zhao, C. Lan, J. Zhou, Isotropic mie resonance-based metamaterial perfect absorber, Appl. Phys. Lett. 103, 031910 (2013) [CrossRef] [Google Scholar]
  18. X. Liu, K. Bi, B. Li, Q. Zhao, J. Zhou, Metamaterial perfect absorber based on artificial dielectric “atoms”, Opt. Express 24, 20454 (2016) [CrossRef] [Google Scholar]
  19. P. Moitra, Y. Yang, Z. Anderson, I.I. Kravchenko, D.P. Briggs, J. Valentine, Realization of an all-dielectric zero-index optical metamaterial, Nat. Photon. 7, 791 (2013) [CrossRef] [Google Scholar]
  20. S. Liu, M.B. Sinclair, T.S. Mahony, Y.C. Jun, S. Cam-pione, J. Ginn, D.A. Bender, J.R. Wendt, J.F. Ihlefeld, P.G. Clem, J.B. Wright, I. Brener, Optical magnetic mirrors without metals, Optica 1, 250 (2014) [CrossRef] [Google Scholar]
  21. B. Luk'yanchuk, N.I. Zheludev, S.A. Maier, N.J. Ha-las, P. Nordlander, H. Giessen, C.T. Chong, The fano resonance in plasmonic nanostructures and metama-terials, Nat. Mater. 9, 707 (2010). [CrossRef] [PubMed] [Google Scholar]
  22. T. Lepetit, E. Akmansoy, J.-P. Ganne, J.-M. Lour-tioz, Resonance continuum coupling in high-permittivity dielectric metamaterials, Phys. Rev. B 82, 195307 (2010) [CrossRef] [Google Scholar]
  23. Q. Zhao, L. Kang, B. Du, H. Zhao, Q. Xie, X. Huang, B. Li, J. Zhou, L. Li, Experimental demonstration of isotropic negative permeability in a three-dimensional dielectric composite, Phys. Rev. Lett. 101, 027402 (2008) [CrossRef] [Google Scholar]
  24. T. Lepetit, E. Akmansoy, M. Pate, J.P. Ganne, Broadband negative magnetism from all-dielectric meta-material, Electron. Lett. 44, 1119 (2008) [CrossRef] [Google Scholar]
  25. I. Vendik, M. Odit, D. Kozlov, 3D isotropic metama-terial based on a regular array of resonant dielectric spherical inclusions, Metamaterials 3, 140 (2009) [CrossRef] [Google Scholar]
  26. J. Kim, A. Gopinath, Simulation of a metamaterial containing cubic high dielectric resonators, Phys. Rev. B 76, 115126 (2007) [CrossRef] [Google Scholar]
  27. T. Lepetit, E. Akmansoy, J.-P. Ganne, Experimental measurement of negative index in an all-dielectric meta-material, Appl. Phys. Lett. 95, 121101 (2009) [CrossRef] [Google Scholar]
  28. H. Nemec, P. Kuzel, F. Kadlec, C. Kadlec, R. Yahiaoui, P. Mounaix, Tunable terahertz metamaterials with negative permeability, Phys. Rev. B 79, 241108 (2009) [CrossRef] [Google Scholar]
  29. M.S. Wheeler, J.S. Aitchison, J.I.L. Chen, G.A. Ozin, M. Mojahedi, Infrared magnetic response in a random silicon carbide micropowder, Phys. Rev. B 79, 073103 (2009) [CrossRef] [Google Scholar]
  30. J.C. Ginn, I. Brener, D.W. Peters, J.R. Wendt, J.O. Stevens, P.F. Hines, L.I. Basilio, L.K. Warne, J.F. Ih-lefeld, P.G. Clem, M.B. Sinclair, Realizing optical magnetism from dielectric metamaterials, Phys. Rev. Lett. 108, 097402 (2012) [CrossRef] [Google Scholar]
  31. J.A. Schuller, R. Zia, T. Taubner, M.L. Brongersma, Dielectric metamaterials based on electric and magnetic resonances of silicon carbide particles, Phys. Rev. Lett. 99, 107401 (2007) [CrossRef] [PubMed] [Google Scholar]
  32. L. Shi, T.U. Tuzer, R. Fenollosa, F. Meseguer, A new dielectric metamaterial building block with a strong magnetic response in the sub-1.5-micrometer region: silicon colloid nanocavities, Adv. Mater. 24, 5934 (2012) [CrossRef] [Google Scholar]
  33. N.I. Zheludev, Y.S. Kivshar, From metamaterials to metadevices, Nat. Mater. 11, 917 (2012) [CrossRef] [Google Scholar]
  34. F. Gaufillet, S. Marcellin, E. Akmansoy, Dielectric metamaterial-based gradient index lens in the terahertz frequency range, IEEE J. Sel. Top. Quantum Electron. 23, 1 (2017) [CrossRef] [Google Scholar]
  35. N. Liu, H. Giessen, Coupling effects in optical meta-materials, Angew. Chem. In t. Ed. 49, 9838 (2010) [CrossRef] [Google Scholar]
  36. F. Zhang, L. Kang, Q. Zhao, J. Zhou, D. Lip-pens, Magnetic and electric coupling effects of dielectric metamaterial, New J. Phys. 14, 033031 (2012) [CrossRef] [Google Scholar]
  37. F. Zhang, V. Sadaune, L. Kang, Q. Zhao, J. Zhou, D. Lippens, Coupling effect for dielectric metamate-rial dimer, Prog. Electromagn. Res. 132, 587 (2012) [CrossRef] [Google Scholar]
  38. S. Lanneb‘ere, S. Campione, A. Aradian, M. Albani, F. Capolino, Artificial magnetism at terahertz frequencies from three-dimensional lattices of TiO2 microspheres accounting for spatial dispersion and magnetoelectric coupling, J. Opt. Soc. Am. B 31, 1078 (2014) [CrossRef] [Google Scholar]
  39. P. Albella, M.A. Poyli, M.K. Schmidt, S.A. Maier, F. Moreno, J.J. Śaenz, J. Aizpurua, Low-loss electric and magnetic field-enhanced spectroscopy with subwave-length silicon dimers, J. Phys. Chem. C 117, 13573 (2013) [CrossRef] [Google Scholar]
  40. S. Marcellin, E. Akmansoy, N, Negative index and mode coupling in all-dielectric metamaterials at terahertz frequencies, in 2015 9th International Congress on Advanced Electromagnetic Materials in Microwaves and Optics (Metamaterials), Oxford, September 2015, pp. 4–6 [CrossRef] [Google Scholar]
  41. T. Lepetit, E. Akmansoy, J.-P. Ganne, Experimental evidence of resonant effective permittivity in a dielectric metamaterial, J. Appl. Phys. 109, 023115 (2011) [CrossRef] [Google Scholar]
  42. N. Matsumoto, T. Hosokura, K. Kageyama, H. Takagi, Y. Sakabe, M. Hangyo, Analysis of dielectric response of TiO2 in terahertz frequency region by general harmonic oscillator model, Jpn. J. Appl. Phys. 47, 7725 (2008) [CrossRef] [Google Scholar]
  43. K. Berdel, J. Rivas, P. Bolivar, P. de Maagt, H. Kurz, Temperature dependence of the permittivity and loss tangent of high-permittivity materials at terahertz frequencies, IEEE Trans. Microw. Theory Tech. 53, 1266 (2005) [CrossRef] [Google Scholar]
  44. R.A. Depine, A. Lakhtakia, A new condition to identify isotropic dielectric-magnetic materials displaying negative phase velocity, Microw. Opt. Tech. Lett. 41, 315 (2004) [CrossRef] [Google Scholar]
  45. S.B. Cohn, Microwave bandpass filters containing high-q dielectric resonators, IEEE Trans. Microw. Theory Tech. 16, 218 (1968) [CrossRef] [Google Scholar]
  46. J. Sethares, S. Naumann, Design of microwave dielectric resonators, IEEE Trans. Microw. Theory Tech. MT14, 2 (1966) [CrossRef] [Google Scholar]
  47. X. Chen, T.M. Grzegorczyk, B.-I. Wu, J. Pacheco, J.A. Kong, Robust method to retrieve the constitutive effective parameters of metamaterials, Phys. Rev. E 70, 016608 (2004) [CrossRef] [Google Scholar]
  48. E. Prodan, C. Radloff, N.J. Halas, P. Nordlander, A hybridization model for the plasmon response of complex nanostructures, Science 302, 5644 419 (2003) [CrossRef] [PubMed] [Google Scholar]
  49. H. Guo, N. Liu, L. Fu, T.P. Meyrath, T. Zentgraf, H. Schweizer, H. Giessen, Resonance hybridization in double split-ring resonator metamaterials, Opt. Express 15, 12095 (2007) [CrossRef] [Google Scholar]
  50. R. Abdeddaim, A. Ourir, J. de Rosny, Realizing a negative index metamaterial by controlling hybridization of trapped modes, Phys. Rev. B 83, 033101 (2011) [CrossRef] [Google Scholar]
  51. N.-H. Shen, L. Zhang, T. Koschny, B. Dastmalchi, M. Kafesaki, C.M. Soukoulis, Discontinuous design of negative index metamaterials based on mode hybridization, Appl. Phys. Lett. 101, 081913 (2012) [CrossRef] [Google Scholar]
  52. N. Liu, H. Guo, L. Fu, S. Kaiser, H. Schweizer, H. Giessen, Plasmon hybridization in stacked cut-wire metamaterials, Adv. Mater. 19, 3628 (2007) [CrossRef] [Google Scholar]
  53. A. Christ, O.J.F. Martin, Y. Ekinci, N.A. Gippius, S.G. Tikhodeev, Symmetry breaking in a plasmonic metamaterial at optical wavelength, Nano Lett. 8, 2171 (2008) [CrossRef] [Google Scholar]
  54. B. Kant́e, Y.-S. Park, K. O’Brien, D. Shuldman, N.D. Lanzillotti-Kimura, Z. Jing Wong, X. Yin, X. Zhang, Symmetry breaking and optical negative index of closed nanorings, Nat. Commun. 3, 1180 (2012) [CrossRef] [Google Scholar]
  55. P. Nordlander, C. Oubre, E. Prodan, K. Li, M.I. Stockman, Plasmon hybridization in nanoparticle dimers, Nano Lett. 4, 899 (2004) [CrossRef] [Google Scholar]
  56. J. van de Groep, T. Coenen, S.A. Mann, A. Polman, Direct imaging of hybridized eigenmodes in coupled silicon nanoparticles, Optica 3, 93 (2016) [CrossRef] [Google Scholar]
  57. A.M. Nicolson, G.F. Ross, Measurement of the intrinsic properties of materials by time-domain techniques, IEEE Trans. Instrum. Meas. 19, 377 (1970) [CrossRef] [Google Scholar]
  58. W. Weir, Automatic measurement of complex dielectric constant and permeability at microwave frequencies, Proc. IEEE 62, 33 (1974) [Google Scholar]
  59. Z. Szabo, G.H. Park, R. Hedge, E.P. Li, A unique extraction of metamaterial parameters based on Kramers-Kronig relationship, IEEE Trans. Microw. Theory Tech. 58, 2646 (2010) [CrossRef] [Google Scholar]
  60. V. Varadan, R. Ro, Unique retrieval of complex permittivity and permeability of dispersive materials from reflection and transmitted fields by enforcing causality, IEEE Trans. Microw. Theory Tech. 55, 2224 (2007) [CrossRef] [Google Scholar]
  61. D.R. Smith, S. Schultz, P. Markǒs, C.M. Soukoulis, Determination of effective permittivity and permeability of metamaterials from reflection and transmission coefficients, Phys. Rev. B 65, 195104 (2002) [CrossRef] [Google Scholar]
  62. T. Koschny, P. Markǒs, D.R. Smith, C.M. Souk-oulis, Resonant and antiresonant frequency dependence of the effective parameters of metamaterials, Phys. Rev. E 68, 065602 (2003) [CrossRef] [Google Scholar]
  63. D.R. Smith, D.C. Vier, T. Koschny, C.M. Souk-oulis, Electromagnetic parameter retrieval from inhomo-geneous metamaterials, Phys. Rev. E 71, 036617 (2005) [CrossRef] [Google Scholar]
  64. A. Alú, Restoring the physical meaning of metamaterial constitutive parameters, Phys. Rev. B 83, 081102 (2011) [CrossRef] [Google Scholar]
  65. C.R. Simovski, On electromagnetic characterization and homogenization of nanostructured metamaterials, J. Opt. 13, 013001 (2011) [CrossRef] [Google Scholar]
  66. R.W. Ziolkowski, Propagation in and scattering from a matched metamaterial having a zero index of refraction, Phys. Rev. E 70, 046608 (2004) [CrossRef] [Google Scholar]
  67. A. Alú, M.G. Silveirinha, A. Salandrino, N. En-gheta, Epsilon-near-zero metamaterials and electromagnetic sources: tailoring the radiation phase pattern, Phys. Rev. B 75, 155410 (2007) [CrossRef] [Google Scholar]
  68. A. Sihvola, Metamaterials and depolarization factors, Prog. Electromagn. Res. 51, pp. 65–82, (2005) [CrossRef] [Google Scholar]
  69. W.G. Spitzer, R.C. Miller, D.A. Kleinman, L.E. Howarth, Far infrared dielectric dispersion in BaTiO3, SrTiO3, and TiO2, Phys. Rev. 126, 1710 (1962) [Google Scholar]
  70. A.S. Barker, Temperature dependence of the transverse and longitudinal optic mode frequencies and charges in srtio3 and BaTiO3, Phys. Rev. 145, 391 (1966) [CrossRef] [Google Scholar]
  71. J. Han, F. Wan, Z. Zhu, W. Zhang, Dielectric response of soft mode in ferroelectric SrTiO3, Appl. Phys. Lett. 90, 031104 (2007) [CrossRef] [Google Scholar]
  72. N. Matsumoto, T. Fujii, K. Kageyama, H. Takagi, T. Na-gashima, M. Hangyo, Measurement of the soft-mode dispersion in SrTiO3 by terahertz time-domain spectroscopic ellipsometry, Jpn. J. Appl. Phys. 48, 09KC11 (2009) [Google Scholar]
  73. F.M.C. Gervais, B. Piriou, Temperature dependence of transverse- and longitudinal-optic modes in TiO2 (rutile), Phys. Rev. B 10, 1642 (1974) [NASA ADS] [CrossRef] [Google Scholar]
  74. T. Tsurumi, J. Li, T. Hoshina, H. Kakemoto, M. Nakada, J. Akedo, Ultrawide range dielectric spectroscopy of BaTiO3-based perovskite dielectrics, Appl. Phys. Lett. 91, 182905 (2007) [CrossRef] [Google Scholar]
  75. J. Petzelt, T. Ostapchuk, I. Gregora, I. Rychetsḱy, S. Hoffmann-Eifert, A.V. Pronin, Y. Yuzyuk, B.P. Gor-shunov, S. Kamba, V. Bovtun, J. Pokorný, M. Savinov, V. Porokhonskyy, D. Rafaja, P. Vaňek, A. Almeida, M.R. Chaves, A.A. Volkov, M. Dressel, R. Waser, Dielectric, infrared, and Raman response of undoped SrTiO3 ceramics: evidence of polar grain boundaries, Phys. Rev. B 64, 184111 (2001) [CrossRef] [Google Scholar]

Cite this article as: Eric Akmansoy, Simon Marcellin, Negative index and mode coupling in all-dielectric metamaterials at terahertz frequencies, EPJ Appl. Metamat. 5, 10 (2018)

All Figures

thumbnail Fig. 1

Schematic layout of the 2D ADM (cross-sectional view). The ADM consists of one infinite layer along the x direction and is made up of two interleaved sets of high permittivity square cross-section dielectric cylinders, which resonate in the first two modes of Mie resonances: the small set resonates in the magnetic mode and the second one in the electric mode. The equidistant cylinders are infinite along the y direction and their side lengths are am = 60 μm and ae = 90 μm, respectively. Their relative permittivity is er = 94. The unit cell actually consists of two subwavelength distinct building blocks. The lattice period is lp= 260 μm and p2 is half the lattice period. The incident wave is transverse electric (TE).

In the text
thumbnail Fig. 2

Spatial mode coupling: minimal value of the effective refractive index neff as a function of the distance p2 between the two resonators. The side lengths of the resonators are am = 60 μm and ae = 90 μm, respectively. Insets: real (purple) and imaginary part (green) of the effective index neff; the shaded area denotes the bandwidth of negative effective index.

In the text
thumbnail Fig. 3

Frequency mode coupling: minimal value of the effective refractive index neff as a function of the side length ae of the electric resonator, namely, the frequency of the electric mode is varying. Insets: real (purple) and imaginary part of (green) the effective index neff; the shaded area denotes the bandwidth of negative effective index. f11 and f12 are the resonances frequencies of the individual resonators (see the text).

In the text
thumbnail Fig. 4

Magnetic mode coupling: frequency (fmr) of the first mode of Mie resonances and of the maximum of absorption (Amr) as a function of the distance p1 between two resonators. The unit cell only comprises the magnetic block. The side length of the resonator is am = 60 μm. The dashed line denotes the resonance frequency (f11) of the mode of resonances of the individual resonator.

In the text
thumbnail Fig. 5

Electric mode coupling: frequency (fer) of the second mode of Mie resonances and of the maximum of absorption (Aer) as a function of the distance p1 between two resonators. The unit cell only comprises the electric block. The side length of the resonator is ae = 90 μm. The dashed line denotes the resonance frequency (f12) of the mode of resonances of the individual resonator.

In the text
thumbnail Fig. 6

Spatial mode coupling: frequency of the first two modes of Mie resonances as a function of the distance p2 between two resonators which is half the lattice period lp. Blue and green colors denote the magnetic and the electric modes, respectively. Square dots and crossed dots denote the maxima of absorption (Amr, Aer) and the minimum of the S12 parameter (fmr, fer), respectively. The shaded area corresponds to negative value of the effective index neff. The side lengths of both resonators are am = 60 μm and ae = 90 μm, respectively. The dashed lines denote the frequencies of resonances (f11, f12) of the modes of the individual resonator.

In the text
thumbnail Fig. 7

Frequency mode coupling: frequency of the first two modes of Mie resonances (fmr, fer) as a function of the side length ae of the electric resonator, namely, the frequency of the electric mode is varying. The convention is the same as in Figure 6. The side length of the magnetic resonator is am = 60 μm and the lattice period lp is 260 μm.

In the text
thumbnail Fig. 8

Effect of spatial mode coupling: S12 parameter (log. scale (dB)) and absorption A (linear scale (lin.)) for several values of distance p2 between two resonators. The side lengths of the resonators are am = 60 μm and ae = 90 μm. (a) Out of frequency degeneracy regime (125 ≤ p2 ≤ 200 μm); (b) in frequency degeneracy regime (100 ≤ p2 ≤ 124 μm). The merged minima of the S12 parameter attain very weak values (≲ − 51 dB) when p2 = p2c = 124 μm, which correspond to the minimum of the refractive index neffmin = −2.2.

In the text
thumbnail Fig. 9

Effect of spatial mode coupling: effective electromagnetic parameters (real and imaginary parts): permeability μeff (top), permittivity eeff (middle) and effective index neff (bottom). (a) Low mode coupling: lattice period lp = 360 μm: the minimum value of the effective refractive index is neffmin < 0.04; (b) strong mode coupling: lattice period lp = 260 μm. The shaded area denotes the bandwidth of the negative effective index: the minimum value of the effective refractive index is neffmin = −1.5. The static effective permittivity values yielded by the Maxwell–Garnett formula, eeff = 2.1 and eeff = 2.9, are in good agreement with the result of the simulation.

In the text
thumbnail Fig. 10

Effect of both mode couplings (spatial mode coupling (SMC) and frequency mode coupling (FMC)): minimal value of the effective refractive index neff as a function of the distance p2 between two resonators for several values of the side length ae of the electric resonator. The side length of the magnetic resonator is am = 60 μm. The two arrows denote increasing mode coupling. The minimum of the real part of the effective index is neffmin = −2.8.

In the text

Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.

Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.

Initial download of the metrics may take a while.